pea protein microgel particles as pickering stabilisers of oil-in ...20revised...111 strong research...

30
This is a repository copy of Pea protein microgel particles as Pickering stabilisers of oil-in- water emulsions: Responsiveness to pH and ionic strength. White Rose Research Online URL for this paper: http://eprints.whiterose.ac.uk/154927/ Version: Accepted Version Article: Zhang, S, Holmes, M, Ettelaie, R et al. (1 more author) (2020) Pea protein microgel particles as Pickering stabilisers of oil-in-water emulsions: Responsiveness to pH and ionic strength. Food Hydrocolloids, 102. ISSN 0268-005X https://doi.org/10.1016/j.foodhyd.2019.105583 © 2019, Elsevier. This manuscript version is made available under the CC-BY-NC-ND 4.0 license http://creativecommons.org/licenses/by-nc-nd/4.0/. [email protected] https://eprints.whiterose.ac.uk/ Reuse This article is distributed under the terms of the Creative Commons Attribution-NonCommercial-NoDerivs (CC BY-NC-ND) licence. This licence only allows you to download this work and share it with others as long as you credit the authors, but you can’t change the article in any way or use it commercially. More information and the full terms of the licence here: https://creativecommons.org/licenses/ Takedown If you consider content in White Rose Research Online to be in breach of UK law, please notify us by emailing [email protected] including the URL of the record and the reason for the withdrawal request.

Upload: others

Post on 31-Jan-2021

3 views

Category:

Documents


0 download

TRANSCRIPT

  • This is a repository copy of Pea protein microgel particles as Pickering stabilisers of oil-in-water emulsions: Responsiveness to pH and ionic strength.

    White Rose Research Online URL for this paper:http://eprints.whiterose.ac.uk/154927/

    Version: Accepted Version

    Article:

    Zhang, S, Holmes, M, Ettelaie, R et al. (1 more author) (2020) Pea protein microgel particles as Pickering stabilisers of oil-in-water emulsions: Responsiveness to pH and ionicstrength. Food Hydrocolloids, 102. ISSN 0268-005X

    https://doi.org/10.1016/j.foodhyd.2019.105583

    © 2019, Elsevier. This manuscript version is made available under the CC-BY-NC-ND 4.0 license http://creativecommons.org/licenses/by-nc-nd/4.0/.

    [email protected]://eprints.whiterose.ac.uk/

    Reuse

    This article is distributed under the terms of the Creative Commons Attribution-NonCommercial-NoDerivs (CC BY-NC-ND) licence. This licence only allows you to download this work and share it with others as long as you credit the authors, but you can’t change the article in any way or use it commercially. More information and the full terms of the licence here: https://creativecommons.org/licenses/

    Takedown

    If you consider content in White Rose Research Online to be in breach of UK law, please notify us by emailing [email protected] including the URL of the record and the reason for the withdrawal request.

  • 1

    1

    Pea protein microgel particles as Pickering stabilisers of 2

    oil-in-water emulsions: Responsiveness to pH and ionic 3

    strength 4

    Shuning Zhang, Melvin Holmes, Rammile Ettelaie, Anwesha Sarkar* 5

    Food Colloids and Bioprocessing Group, School of Food Science and Nutrition, University of 6

    Leeds, Leeds, LS2 9JT, UK 7

    8

    9

    10

    11

    12

    13

    14

    15

    *Corresponding author: 16

    Dr. Anwesha Sarkar 17

    Food Colloids and Bioprocessing Group, 18

    School of Food Science and Nutrition, University of Leeds, Leeds LS2 9JT, UK. 19

    E-mail address: [email protected] (A. Sarkar). 20

    21

    mailto:[email protected]

  • 2

    Abstract 22

    The aim of this study was to design plant protein-based microgel particles to create Pickering 23

    emulsions (20 wt% sunflower oil, 0.05-1.0 wt% protein) and investigate the role of electrostatic 24

    interactions on colloidal behaviour of such emulsions. Pea protein microgels (PPM) were 25

    designed using a facile top-down approach of heat-set protein gel formation followed by 26

    controlled shearing. The aqueous dispersion of PPM had hydrodynamic diameters ranging 27

    from 200 to 400 nm at pH 7.0 to pH 9.0 with high negative charge (-30 to -35 mV) and pI was 28

    pH 5.0. With increasing ionic concentration from 1 to 250 mM NaCl, the ζ-potential of PPM 29

    changed to -8 mV due to charge screening effects, in line with theoretical calculations of the 30

    electrostatic potential. The Pickering emulsions with smallest droplet sizes (d43) ~25 µm 31

    exhibited excellent coalescence stability and high adsorption efficiency of PPM at the oil-water 32

    interface (>98%) at pH 7.0, with the latter being supported by confocal microscopy showing 33

    effective adsorption of the PPMs at the droplet surface. Adjusting the pH of the emulsions to 34

    pI demonstrated aggregation of adsorbed PPM at the particle-laden interface providing a higher 35

    degree of adsorption as well as enhancing inter-droplet flocculation and the shear-thinning 36

    character as compared to those at pH 7.0 or pH 3.0. Charge screening effects in presence of 37

    100 mM NaCl resulted in PPM-PPM aggregation and enhanced viscosity of the emulsions. 38

    Findings from this study on pea protein microgels would open avenues for rational designing 39

    of sustainable Pickering emulsions in the future. 40

    41

    Keywords: Pickering emulsion; pea protein; microgel; interaction potential; plant protein; 42

    particle-stabilized interface 43

    44

    45

  • 3

    1. Introduction 46

    In recent times, gradual dietary changes towards sustainable plant-based ingredients has 47

    increased and, consequently designing food emulsions using plant proteins have been the 48

    preferred direction for emulsifiers to reduce food-production associated environmental 49

    footprints (Rayner, et al., 2014). Furthermore, plant proteins are appreciated by consumers as 50

    they are “green”, “vegan-friendly” and are considered to be less allergenic and are less 51

    expensive as compared to the most commonly used dairy proteins. A range of plant proteins, 52

    such as soybean, pea, chickpea, faba bean, lentil, cowpea, French bean, sweet lupin, tomato 53

    seed and zein protein (Ben-Harb, et al., 2018; Jayasena, et al., 2010; Karaca, et al., 2011; 54

    Kimura, et al., 2008; Pan, et al., 2015; Sarkar, et al., 2016a) have been investigated for 55

    stabilizing oil-in-water (O/W) emulsions. However, many if not most, of these plant proteins 56

    have limited aqueous solubility (Makri, et al., 2006; Nesterenko, et al., 2013; Sarkar, et al., 57

    2016a) and are less digestible as compared to the dairy proteins (de Folter, et al., 2012; Laguna, 58

    et al., 2017). This restricts easy replacement of dairy proteins in food products by plant proteins. 59

    60

    In comparison to the conventional protein-stabilized emulsions, the investigation of plant 61

    protein-based particles to create Pickering emulsions can be particularly interesting as these 62

    emulsions involve particle-stabilization of the droplets and therefore do not need perfect 63

    solubilisation of these proteins in the aqueous phase. In other words, these emulsions need the 64

    particles to be partially wettable by the oil and aqueous phases. In addition, Pickering 65

    emulsions have gained remarkable research interest in the food colloids community in recent 66

    years due to their distinctive stability against coalescence and Ostwald ripening as well as their 67

    ability to alter lipid digestion kinetics of emulsions post consumption (Aveyard, et al., 2003; 68

    Binks, 2002; Dickinson, 2012; Dickinson, 2013; Sarkar, et al., 2019). Hence, there is a strong 69

  • 4

    demand in food industries to find cheaper plant protein-based Pickering stabilizer alternatives 70

    that can effectively stabilize emulsion droplets for longer term against coalescence. 71

    72

    As compared to the extensive studies on particles derived from animal-based proteins being 73

    used as Pickering stabilizers, e.g. whey protein microgels (Destribats, et al., 2014; Sarkar, et 74

    al., 2016b) and lactoferrin nanoparticles or nanogels (David-Birman, et al., 2013; Gal, et al., 75

    2013; Meshulam, et al., 2014; Sarkar, et al., 2018), those involving particles obtained from 76

    plant proteins are fairly recent and are attracting significant research attention. For example, 77

    Liu and co-authors (Liu, et al., 2017; Liu, et al., 2013; Liu, et al., 2014; Liu, et al., 2016; Peng, 78

    et al., 2020; Peng, et al., 2018; Zhu, et al., 2017) investigated the ability of soy protein 79

    nanoparticles aggregates (SPN) (~ 100 nm, created via heat-treatment at 95 oC for 15 min) to 80

    act as Pickering stabilizers. On the other hand, Chen, et al. (2014) prepared heat denatured soy 81

    protein nanogel particles at various pH values (pH 2.0-7.0) and added ions (0-200 mM NaCl). 82

    In another recent study, Zhu, et al. (2018) suggested the importance of the electrostatic 83

    screening by ions (100-200 mM NaCl) to improve freeze-thaw stability of Pickering emulsions 84

    stabilized by soy protein-based nanoparticles. Besides the commonly used soy protein-based 85

    particles, peanut protein microgel particles have been recently investigated, where these 86

    microgel particles were prepared via enzyme treatment and had hydrodynamic diameters 87

    ranging from 200 to 300 nm and were used to stabilize high-internal-phase Pickering emulsions 88

    with 87 % oil volume fractions (Jiao, et al., 2018). Water-insoluble zein-based colloidal 89

    particles and kafirin nanoparticles have also been reported as possible Pickering stabilizers. For 90

    example, de Folter, et al. (2012) fabricated zein-based colloidal particles with an average 91

    diameter of ~ 70 nm. Xiao, et al. (2016) used an anti-solvent precipitation method to produce 92

    kafirin nanoparticles with a mean diameter of 206.5 nm. Gliadin colloidal particles (GCPs) 93

  • 5

    with average diameter of ~120 nm at acidic pH that were prepared by an anti-solvent procedure 94

    (Hu, et al., 2016) were suggested as Pickering stabilisers. 95

    96

    Due to the significant academic and industrial interests resulting from their low cost, Liang, et 97

    al. (2014) created pea protein nanoparticles for the first time. In such a process, an aqueous 98

    dispersion of pea protein isolate (PPI) was adjusted at pH 3.0 to produce pea protein-based 99

    particles with hydrodynamic diameter of 100-200 nm. In addition, such particles generated oil-100

    in-water (O/W) Pickering emulsion with 20 wt% oil volume faction. In another study, Shao, et 101

    al. (2015) found that such pea protein particle-stabilized Pickering emulsion enabled controlled 102

    release of lipophilic bioactive compounds during in vitro gastrointestinal digestion. Compared 103

    to the emulsion stabilized with untreated PPI, the Pickering-stabilised emulsion had a sustained 104

    release behaviour due to their gel-like inter-droplet network formation. Another recent study 105

    by Cochereau, et al. (2019) designed pea protein microgel particles with protein concentration 106

    of 1-4 wt% at pH 6-8 via slow and modest heat treatment (i.e. 20- 40 ℃). 107

    108

    It is thus clear from the literature that stabilizing Pickering emulsions using plant 109

    protein-based particles is a relatively recent endeavour. In particular, considering the 110

    strong research interests by food industries and academic community in pea protein, it 111

    is surprising that relatively little attention has been devoted to designing pea protein-112

    based particles for the purpose of stabilizing Pickering emulsion droplets. Although the 113

    gelation properties of pea proteins (Bora, et al., 1994; Mession, et al., 2015; Shao, et al., 114

    2015), pea protein-based aggregates, such as heat-treated fibrillar aggregates (Munialo 115

    et al., 2014), mixed pea globulin aggregates (Mession, et al., 2017), as well as thermal 116

    aggregates from mixed pea globulin and β-lactoglobulin (Chihi, et al., 2016; Chihi, et 117

    al., 2018), have been widely studied, there has only been two studies that have used pea 118

  • 6

    protein particles to prepare Pickering emulsions (Liang, et al., 2014; Shao, et al., 2015). 119

    Furthermore, these two studies have been performed using pea protein gelation at just 120

    one particular pH (pH 3.0), thus restricting the use of such emulsions over a wider range 121

    of pH and ionic strengths. To our knowledge, there has been no study that has 122

    systematically characterized the colloidal properties of thermally-crosslinked pea 123

    protein microgel in the aqueous phase, as well as when present at the droplet surface, 124

    and in particular also investigated the role of electrostatics on the colloidal behaviour of 125

    these types of emulsions. Considering the recent demand of plant-based protein 126

    particles, it is necessary to characterise the ability of such Pickering emulsions at a wide 127

    range of pH and ionic strengths to understand their responsiveness to environmental 128

    conditions during their processing and indeed after consumption. 129

    130

    Hence, in this study, pea protein has been used to design pea protein microgels (PPM) 131

    using a top down approach for creating a heat-set hydrogel, followed by controlled 132

    shearing to investigate their potential to stabilize O/W Pickering emulsions, which has 133

    not been reported in literature to date. We hypothesized that pea protein microgel 134

    already adsorbed to the interface would aggregate at the droplet surface by suitably 135

    adjusting the pH to acidic pH, forming a densely packed layer of particles further 136

    protecting the oil droplets against coalescence. Colloidal stability of pea protein 137

    microgel particles (PPM) in aqueous phase and PPM-stabilized emulsions were 138

    systematically characterized as a function of pH (pH 2.0-9.0) and ionic strength (1-250 139

    mM NaCl) using particle sizing (dynamic light scattering), droplet sizing (static light 140

    scattering), optical microscopy, confocal laser scanning microscopy (two dimensional 141

    (2D) as well as three dimensional (3D) images), apparent viscosity, adsorption 142

    efficiency assessment and ζ-potential measurements. The composition of PPM at the 143

  • 7

    interface was assessed using sodium dodecyl sulphate polyacrylamide gel 144

    electrophoresis (SDS-PAGE). In addition, we calculated the interaction potentials of the 145

    particles as a function of pH and ionic strengths at close separation distances to identify 146

    the electrostatic contribution of these particles at the droplet surface during any droplet 147

    aggregation phenomena. 148

    149

    2. Materials and Methods 150

    2.1 Materials 151

    Commercial pea protein concentrate (Nutralys S85XF) (PPC) with 85% protein content was 152

    kindly gifted by Roquette (Lestrem, France). Sunflower oil was purchased from local 153

    supermarket and used without any further purification. Mini-PROTEAN TGX Gels, ProtoBlue 154

    Safe Colloidal Coomassie G-250 stain and all sodium dodecyl sulphate polyacrylamide gel 155

    electrophoresis (SDS-PAGE) chemicals were purchased from Bio-Rad Laboratories, UK. 156

    Sodium azide, Nile Red and Nile Blue were purchased from Sigma Aldrich (Dorset, UK). All 157

    other chemicals were of analytical grade and purchased from Sigma-Aldrich Dorset, UK) 158

    unless otherwise specified. All solutions were prepared with Milli-Q water (water purified by 159

    a Milli-Q apparatus, Millipore Corp., Bedford, MA, USA) with a resistivity of 18.2 MΩ cm at 160

    25 ℃. 161 162

    2.2 Methods 163

    2.2.1 Preparation of pea protein microgel (PPM) 164

    Pea protein-based Pickering particles were prepared using slight modification of the top-down 165

    method developed by Sarkar, et al. (2016b). Briefly, the PPC powder at 15 wt% powder i.e. 166

    12.75 wt% protein was dispersed in 20 mM phosphate buffer at pH 7.0. The aqueous dispersion 167

  • 8

    of PPC was mixed for 2 hours using magnetic stirring at 300 rpm, and then stored at 4 oC 168

    overnight. The PPC dispersion was heated at 90 oC for 1 hour to allow formation of heat-set 169

    gel (Figure 1). During heat treatment, the globular pea proteins were denatured and unfolded 170

    (Laguna, et al., 2017). And then the denatured proteins aggregated via the disulphide 171

    crosslinking forming a macroscopic protein-based hydrogel. After cooling to room temperature 172

    using flowing water at 25 oC, the pea protein hydrogel was stored at 4 oC overnight. These 173

    hydrogels were mixed with 20 mM phosphate buffer (1: 1 v/v) at pH 7.0 and were broken to 174

    macrogel particles using a blender (HB711M, Kenwood, UK) at speed 3 for 5 minutes. After 175

    removing the air bubbles generated during blending using vacuum (25-30 bar) for 15 min, the 176

    macroscopic gel particle dispersion was homogenized using two passes through a two-stage 177

    valve homogenizer (Panda Plus 2000, GEA Niro Soavi Homogeneizador Parma, Italy) 178

    operating at first/ second stage pressures of 250/ 50 bars, respectively. The resultant particles, 179

    termed as pea protein microgels (PPM) contained 6.375 wt% protein. 180

    181

    2.2.2 Preparation of PPM-stabilized Pickering emulsions (PPM-E) 182

    Sunflower oil was mixed with appropriate quantities of PPM at 20:80 oil: protein (w/w) ratio 183

    using rotor-stator (L5M-A, Silverson machines, UK) mixing at 8,000 rpm for 5 minutes. The 184

    PPM dispersion was diluted using phosphate buffer to have 0.05, 0.10, 0.25, 0.50 and 1.0 wt% 185

    protein content in the final emulsions, henceforth, such emulsions are referred as E0.05, E0.1, 186

    E0.25, E0.5 and E1.0, respectively. The pre-homogenized oil-PPM mixture was further 187

    homogenized by a two-stage valve homogenizer (Panda Plus 2000, GEA Niro Soavi 188

    Homogeneizador Parma, Italy) operating at two stages, of 250 and 50 bar pressures (Figure 1) 189

    resulting in Pickering emulsions (PPM-E, E0.05-E1.0). Sodium azide (0.02 wt%) was added 190

    as an antimicrobial agent. 191

  • 9

    2.2.3 Coalescence stability of PPM-E during storage 192

    The PPM-Es (E0.05, E0.10, E0.25, E0.50 and E1.0) were stored at 4 oC for a period of 28 days 193

    and were monitored using droplet sizing, ζ-potential measurement and visual observation to 194

    select the most stable emulsions for pH and ion treatment. 195

    196

    2.2.4 pH and ion treatment 197

    To investigate the colloidal stability of PPM and PPM-E (E1.0, chosen based on the 198

    coalescence stability study) under environmental conditions, the samples were subjected to 199

    different pH and ionic strengths. The pH adjustment (pH 2.0-9.0) was done by both “low to 200

    high” and “high to low” methods by adding 1 M HCl or 1 M NaOH, without any salt addition 201

    (Adal, et al., 2017). For “high to low” pH adjustment, the PPM at pH 9.0 was rapidly adjusted 202

    to a target pH while mixing. For “low to high”, the PPM at pH 2.0 was adjusted to a target pH 203

    quickly while mixing. In case of E1.0, the pH responsiveness was checked at pH 3.0, 5.0 and 204

    7.0. For ionic strengths, the pH value of PPM was kept constant at pH 7.0 and ionic strength 205

    was adjusted from 1-250 mM NaCl. For E1.0, the physicochemical stability was assessed by 206

    subjecting the emulsions to 1 mM, 10 mM and 100 mM NaCl, respectively. 207

    208

    2.2.5 Size and ζ-potential measurements. 209

    The hydrodynamic diameters of PPM dispersion at various pH and ionic strengths were 210

    measured at 25 oC using a Zetasizer Nano-ZS (Malvern Instruments Ltd, Malvern, 211

    Worcestershire, UK). The PPM sample was diluted to 0.004 wt% particle concentration with 212

    Milli-Q water before measurement. Assuming that there is no particle–particle interaction in 213

    the diluted sample, the hydrodynamic diameter (Dh) of the droplets was calculated using the 214

    Stokes–Einstein equation (Equation 1): 215

  • 10

    𝐷ℎ = 𝑘𝐵𝑇3𝜋𝜂𝐷𝑡 (1) 216 where Dt is the translational diffusion coefficient, kB is Boltzmann’s constant, T is 217

    thermodynamic temperature, and η is dynamic viscosity. The refractive index of PPM was set 218

    at 1.52. The absorbance of the protein was assumed to be 0.001. 219

    Droplet size distribution of PPM-Es (E0.05, E0.10, E0.25, E0.50 and E1.0) as a function of 220

    storage and E1.0 as a function of different pH or with various ionic strengths were determined 221

    using static light scattering (SLS) techniques using a Malvern MasterSizer 3000 (Malvern 222

    Instruments Ltd, Malvern, Worcestershire, UK) at 25 ℃. Samples were added dropwise in 223 distilled water until the instrument gave an obscuration rate between 4 and 6%. The average 224

    droplet size of the emulsion was reported as the volume mean diameter (d43= ∑ 𝑛𝑖𝑑𝑖4∑ 𝑛𝑖𝑑𝑖3 ) and 225 surface mean (d32= ∑ 𝑛𝑖𝑑𝑖3∑ 𝑛𝑖𝑑𝑖2) 226 where, ni is the number of droplets with a diameter, di. The refractive index of sunflower oil 227

    and the dispersion medium were set at 1.46 and 1.33, respectively. The absorbance value of 228

    the emulsion droplets was set at 0.001. 229

    230

    Zetasizer Nano ZS was used to measure the ζ-potential of PPM dispersion, PPM-Es (E0.05, 231

    E0.10, E0.25, E0.50 and E1.0) as a function of storage and E1.0 as a function of pH and ionic 232

    strength. Before measurement, PPM dispersion at different pH values was diluted to 0.004 wt% 233

    particle concentration and E1.0 with different pH values was diluted to 0.005 wt% droplet 234

    concentration using Milli-Q water adjusted to relevant pH (2.0-9.0). Similarly, PPM or E1.0 235

    containing different concentrations of NaCl was diluted Milli-Q water adjusted to relevant 236

    NaCl concentrations (1-250 mM). The diluted sample was then added into a folded capillary 237

    cell (DTS1070 cell, Malvern Instruments Ltd., Worcestershire, UK), which had two electrodes. 238

  • 11

    After 120 seconds of equilibration in the Zetasizer at 25 oC, the PPM particles or E1.0 droplets 239

    moved towards oppositely charged electrodes. The magnitude of ζ-potential was determined 240

    from the terminal speed of the particle motion using Henry’s equation, with Smoluchowski 241

    approximation appropriate here since the thickness of the diffused double layer is expected to 242

    be much smaller than the size of the particles. 243

    244

    2.2.6 Adsorption efficiency of PPM at the oil-water interface 245

    To determine the adsorption efficiency of PPM at the oil-water interface after pH-treatment 246

    (adjusting the pH of PPM-E to pH 3.0, 5.0 and 7.0) or salt-treatment (100 mM NaCl), the 247

    quantity of PPM in the emulsion phase was determined (Araiza-Calahorra, et al., 2019; Sarkar, 248

    et al., 2016b). Briefly, PPM-E (E1.0) at different pHs (pH 3.0, 5.0 and 7.0) and ionic strength 249

    (100 mM NaCl) were diluted (1:4 w/w) with phosphate buffer (pH 7.0) or Milli-Q water 250

    (adjusted to pH 5.0 or pH 3.0) or phosphate buffer at H 7.0 containing 100 mM NaCl. All 251

    diluted emulsions were centrifuged for 40 mins at 10,000 rpm, 20 oC (Fresco 21 centrifuge, 252

    Thermo Fisher Scientific, Germany). The subnatants were carefully collected using a syringe 253

    and then measured using a DC protein assay kit (Bio-Rad Laboratories, Hercules, CA) on UV-254

    Vis Spectrophotometer with an adsorption wavelength of 750 nm. The adsorption efficiency 255

    of PPM at the interface was calculated by subtracting the amount of PPM in the subnatant from 256

    the total amount of PPM used initially to prepare the emulsions as a percentage of total protein 257

    concentration in the emulsion. 258

    259

    2.2.7 Apparent viscosity measurement 260

    The apparent viscosity of PPM-E (E1.0) as a function of pH and ionic strength were determined 261

    at 25 °C using a Kinexus ultra rheometer (Malvern Instruments Ltd, Malvern, UK). The 262

  • 12

    apparent viscosities (ηa, Pa s) were recorded as a function of shear rates ranging from 0.1 to 263

    1000 s−1. All measurements were done in triplicates and were reported as the mean and standard 264

    deviation. In order to determine the flow type the emulsions as a function of pH and ionic 265

    strength, the flow curves were fitted using power-law model (Ostwald-de Waele model) (see 266

    Equation (2)): 267

    𝜂𝑎(�̇�) = 𝐾(�̇�)𝑛−1 (2) 268 where, K is consistency coefficient (Pa sn), n is power-law index and �̇� is shear rate (s-1). 269 270

    2.2.8 Microscopy 271

    Optical microscopy (Nikon, SMZ-2T, Japan) was used to observe the microstructure of PPM 272

    and E1.0 as a function of pH and ionic strengths. The samples undergoing optical microscopy 273

    needed to be diluted 10 times in respective buffer. Zeiss confocal microscope (Model LSM 274

    700, Carl Zeiss MicroImaging GmbH, Jena, Germany) was used for microstructural 275

    characterization of PPM at the interface of E1.0 droplets. The oil droplets in E1.0 were stained 276

    with 100 μL of Nile Red (2% w/v in in dimethyl sulfoxide) and the protein stabilizing the oil 277

    droplets was stained with 500 μL of Nile Blue (10% w/v in Milli-Q water). Nile Red was 278

    excited by at 488 nm whereas Nile Blue was excited at 635 nm. The stained samples were 279

    mixed with an appropriate amount of xanthan gum (1 wt %) to reduce the Brownian motion of 280

    oil droplets. The prepared samples were placed onto a microscope slide, covered with a cover 281

    slip and observed at 63 × (oil) magnifications. 282

    283

    2.2.9 Sodium dodecyl sulphate-polyacrylamide gel electrophoresis (SDS-PAGE) 284

    Sodium dodecyl sulphate polyacrylamide gel electrophoresis (SDS-PAGE) under reducing 285

    conditions was used to determine the composition of protein in the initial pea protein 286

  • 13

    solution, PPM dispersions and the adsorbed phase of PPM-E (E1.0). The samples included 287

    PPM dispersions that are pH-treated (adjusting the pH of PPM to pH 3.0, 5.0 and 7.0) and 288

    salt-treated (100 mM NaCl) PPM dispersions as well as PPM-E (E1.0) at different pHs (pH 289

    3.0, 5.0 and 7.0) and ionic strength (100 mM NaCl). To determine the protein compositions 290

    of absorbed phase i.e. the particles at the interace, all the PPM-E samples after pH and ionic 291

    treatments were centrifuged at 14,500 g for 45 min, and then the cream phases were 292

    carefully collected, dispersed in Milli-Q water and centrifuged again for 45 min at 14,500 g. 293

    Approximately, 65 μL of pea protein dispersion, PPM samples and cream layer of PPM-E 294

    (E1.0) samples were mixed with 25 μL pf SDS loading buffer (62.5 mM Tris-HCl, pH 6.8, 295

    2% SDS, 25% glycerol, 0.01% bromophenol blue) and 10 μL of diothiothreitol (DTT, of a 296

    final concentration of 50 mM), heated at 95 ℃ for 5 min.The SDS-PAGE was carried out 297 by loading 5 μL of protein marker and 10 μL of these samples-SDS buffer mixtures in the 298

    Mini-PROTEAN 8-10% TGX Gels in a Mini-PROTEAN II electrophoretic unit Bio-Rad 299

    Laboratories, Richmond, CA, USA). The resolving gel contained 16% acrylamide and the 300

    stacking gel was made up of 4% acrylamide. 301

    302

    The running process had two stages; one at 100 V for 10 min at first stage and then 200 V 303

    of about 20 min for the second stage. After the run, the gel was stained for 2 hours using 304

    Coomassie Blue solution, which consisted of 90% ProtoBlue Safe Colloidal Coomassie G-305

    250 stain and 10% ethanol. The gel was then destained using Milli-Q water overnight and 306

    scanned using the ChemiDoc™ XRS+ with image LabTM Software (Bio- Rad Laboratories, 307

    Inc, USA). Each band within the lanes was selected automatically by the software to cover 308

    the whole band. Background intensity was subtracted after scanning an empty lane. The 309

    SDS-PAGE experiments were carried out in triplicates and band intensities was reported 310

    as an average and standard deviation of three reported readings. 311

  • 14

    2.2.10 Statistical analysis 312

    All experimental results were reported as mean and standard deviations of five measurements 313

    on triplicate samples (n = 5 × 3). The statistical analyses were conducted using one-way 314

    ANOVA and multiple comparison test using SPSS software (IBM, SPSS statistics, version 24) 315

    and the significant difference between samples were considered when p < 0.05. 316

    317

    3. Results and Discussion 318

    3.1. Characteristics of aqueous dispersion of PPM 319

    The aqueous dispersion of PPM in phosphate buffer at pH 7.0 had a narrow size distribution 320

    with single peak in the size range of 100 to 1000 nm (Figure 2). The PPM had a hydrodynamic 321

    diameter (Dh) of about 232 nm (polydispersity index (PDI) < 0.2) (Figure 3a). Although an 322

    aqueous dispersion of PPC showed multimodal size distribution with high PDI of nearly 1.0 323

    (data not shown), preparing PPM via the top-down approach of heat-set gel formation and 324

    controlled shearing appears as a feasible approach to create plant protein-based particles with 325

    high colloidal stability. For instance, the PPM were highly negatively charged (-40 mV) 326

    (Figure 3b) showing no particle aggregation or macroscopic sedimentation (Figure 3c) at pH 327

    7.0, respectively over a year storage. 328

    329

    3.2 Colloidal stability of aqueous dispersions of PPM 330

    As shown in Figure 3a, Dh of PPM ranged from 200 to 400 nm at neutral to alkaline pH (pH 331

    7.0 to pH 9.0) (p > 0.05). However, at pH 6.0, PPM showed a marked increase in Dh supported 332

    by correspondingly steep increase in PDI to 0.8 as compared to that at neutral pH (p < 0.05). 333

    The Dh of PPM dispersions increased to the highest values of > 8,000 nm at pH 5.0 followed 334

  • 15

    by a decrease to < 2,000 nm at 3.0 < pH < 5.0 (Figure 3a). High degree of particle aggregation 335

    and sedimentation observed using optical microscopy and macroscopic images, respectively, 336

    indicate pH 5.0 to be the isoelectric point (pI) (Figure 3c). Caution should be exercised while 337

    interpreting these Dh results with values above 1000 nm, when using dynamic light scattering 338

    (Supplementary Figure S1). Thus, focusing on the trend of increasing values of Dh, the 339

    colloidal stability of aqueous dispersions of PPM was limited at pH 6.0 with high PDI (Figure 340

    3a) and excessive particle aggregation (Figure 3c). 341

    342

    The colloidal behaviour of PPM as a function of pH was investigated using ζ-potential 343

    measurements of PPM (Figure 3b). The ζ-potential of PPM increased from -40 mV to +30 mV 344

    with pH change, from pH 9.0 to pH 2.0 (p < 0.05). Noticeably, the net surface charge was close 345

    to zero at pH 5.0 (Figure 3b) validating this to be the pI, corroborating with the largest 346

    hydrodynamic diameter and PDI (almost close to 1.0) data (Figure 3a). At this pH, the 347

    positively-charged amino-groups in PPM were balancing the negatively-charged carboxyl 348

    groups. Interestingly, the pI of PPM (Figure 3b) shifted slightly from that of the PPC, where 349

    latter is reported to be around 4.0 (Adal, et al., 2017). Such discrepancy in pI values between 350

    protein concentrates and protein microgel particles are possible owing to the unfolding process 351

    during thermal treatment of the latter and consequently exposure of some charged groups. At 352

    or below pH 4.0, PPM showed net positive charge ranging from +20 to +30 mV. An interesting 353

    study by Destribats, et al. (2014) demonstrated that in whey protein microgel particles (WPM), 354

    some larger particle were present even when the pH value was far from the pI and the particles 355

    possessed electrostatic charge. The larger particles were postulated to be associated with the 356

    swelling of WPM i.e. the solvation of the exposed protein groups enabling WPM to swell. 357

    Nevertheless, in the present study even if some degree of swelling of PPM might have occurred, 358

  • 16

    due to the solvation of the protein groups that was promoted at acidic pH, such effects have 359

    been overshadowed by the dominant particle aggregation as observed in Figure 3c. 360

    361

    It is worth noting that the PPM in presence of electrolytes showed no significant difference in 362

    Dh and PDI (Figure 4a) as a function of ionic strengths. This was in close agreement with no 363

    aggregation or sedimentation observed in optical microscopy or macroscopic images (Figure 364

    4c). However, a progressive decrease of ζ-potential magnitude from -31 to -7.5 mV was 365

    observed (Figure 4b) as the ionic strength increased from 1 to 250 mM NaCl. Similar salt-366

    induced reduction in ζ-potential in plant protein-based particles has been observed in the case 367

    of soy protein-based nanoparticles, where Liu, et al. (2013) demonstrated that the increasing 368

    NaCl concentration (0-500 mM) led to the decrease of absolute magnitude of ζ-potential of soy 369

    protein nanoparticles. This is due to a decreased Debye length which for constant charged 370

    surfaces at any rate will lead to a lower ζ-potential. 371

    372

    Colloidal particles, such as PPM in this study, dispersed in aqueous solutions will experience 373

    Derjaguin-Landau-Verwey-Overbeek (DLVO) forces, other types of short-ranged, attractive 374

    non-DLVO forces (Hogg, et al., 1966) as well as possible steric repulsion. To understand the 375

    role of interaction in the colloidal stability of PPM, we used DLVO theory to calculate the 376

    inter-particle electrostatic-mediated activation energy barrier between sub-micron sized 377

    PPMs, as two such particles approach each other at varying pHs and ionic strengths (Figure 378

    5). This is done using equation (3) (Cosgrove, 2010): 379

    UR=2πεψ2aln[1+exp(-kh)] (3) 380

    for the electrostatic component of particle-particle interaction. For equation (3), ε is the 381

    permittivity of the system, i.e. ε0εr (ε0 is the permittivity of vacuum, and εr is relative 382

  • 17

    permittivity of water), ψ is the surface potential approximately equal to the ζ-potential values 383

    measured at different pHs or ionic strengths, a is the radius of PPM particle (i.e. 118 nm), h is 384

    the particle-particle surface separation distance ranging from 0.1 to 5 nm and κ-1 is the Debye 385

    length, which was calculated from equation (4): 386

    κ =[𝑁𝐴𝑒2𝜀𝑘𝑇 ∑ 𝑧𝑖2𝑐𝑖∞𝑖 ]12 (4) 387 Here, 𝑧𝑖 is the valency of ith type of ion, 𝑐𝑖∞ is the number density of that ion, NA is the 388 Avogadro’s number, and T is the temperature (298 K). 389

    The total interaction potential (U) between PPMs is the summation of electrostatic repulsion 390

    (UR) and van der Waals attraction (UVW). In this calculation, the pH independent van der Waals 391

    attractive energy (UVW) between two equal sized PPM was calculated using equation (5): 392

    UVW = − 𝑎𝐴𝐻12ℎ (5) 393 where, AH is the Hamaker constant which has been assumed to be 1 kBT, similar to other 394

    reported protein particles (Tuinier, et al., 2002). 395

    The total interaction potential was higher than 100 kBT at pHs < 3.0 and pHs > 6.0 (Figure 5a), 396

    suggesting an electrostatically-induced energy barrier being sufficient to slow down any 397

    aggregation to negligible rates (see Supplementary Figure S2a for electrostatic repulsion). 398

    However, the van der Waal’s attractive forces particularly for PPM in the acidic pH i.e. pH 399

    2.0-4.5 played a dominant role when at close PPM-PPM separation distance (< 0.15 nm). This 400

    caused the total energy maximum to fall below ~10 kBT (Figure 5a), in line with high Dh, high 401

    PDI (Figure 3a) and extensive particle aggregation (Figure 3c). The ζ-potential of PPM 402

    reflected minimal surface charge (almost close to zero, Figure 3b) at pH 5.0 and the 403

    correspondingly lack of sufficient electrostatic repulsion barrier, to provide a large energy in 404

  • 18

    the presence of strong van der Waal’s attraction (Figure 5a). A low energy barrier accelerated 405

    the aggregation of the particles (Figures 3a and 3c). 406

    407

    The UR and UVW between PPMs with varying ionic strengths were also calculated using 408

    equations (3-5), as shown in Figure 5b (see Supplementary Figure S2b for UR and UVW). 409

    Interestingly, at 1-10 mM NaCl, the interaction was mainly dominated by electrostatic 410

    repulsive forces with PPM-PPM interaction potential of > 100 kBT overshadowing effects of 411

    van der Waal’s attractive forces (Figure 5b) over most separation distances. As expected, the 412

    electrostatic repulsion between particles was screened, down to nearly zero, with the increase 413

    in the NaCl concentration (Figure 5b). Once the salt concentration was above 50 mM, the 414

    dominating van der Waal’s attractive forces between PPMs should result in particle 415

    aggregation, leading to an unstable dispersion. However, this was not the case experimentally. 416

    In fact, larger particle aggregates were not observed (Figure 4c), and the size of PPM was 417

    stable at a range of 230 to 240 nm, with lower PDI values (Figures 4a). The possible 418

    explanation for the discrepancy between the theoretically predicted aggregation and an 419

    experientially observed stable dispersion might be attributed to the steric repulsion effects 420

    associated with the hairy PPM particles, which is not strongly influenced by electrolyte 421

    concentration. This requires experimental investigation in the future. This behaviour is unlike 422

    that of WPM, soy protein nanoparticles, zein colloid particles and kafirin nanoparticles (de 423

    Folter, et al., 2012; Destribats, et al., 2014; Liu, et al., 2013; Xiao, et al., 2016), where NaCl 424

    addition is known to cause aggregation of particles in the aqueous dispersions. 425

    426

    3.3 Characteristics of Pickering O/W emulsions stabilized by PPM (PPM-E) 427

    3.3.1 Stability of PPM-E prepared using various concentration of PPM 428

  • 19

    To determine the optimum concentration of PPM to stabilize PPM-E, coalescence stability of 429

    PPM-Es stored under refrigerated conditions for a period of 28 days were characterized using 430

    size, charge and visual imaging of any oiling off. Table 1 shows the emulsion droplet size 431

    and ζ-potential, and Figure 7 presents the corresponding visual images of the freshly prepared 432

    PPM-Es (E0.50-1.0) after 7, 14 and 28 days storage, respectively, at 4 ℃. The cream layer was 433

    evident in all the freshly prepared emulsions (E0.05-0.5), except in E1.0 (Figure 7), however, 434

    all the emulsions reverted back to a homogenous dispersion after gentle hand shaking without 435

    any visual evidence for coalescence. After a week of storage, although no phase separation was 436

    discernible in E0.05 and E0.1 macroscopically (Figure 7), larger oil droplets were visible after 437

    diluting the emulsions with buffer (Supplementary Figure S3) with clear signs of coalescence 438

    and hence the size and ζ-potential could not be measured (Table 1) in these emulsions. This is 439

    expected as the insufficient quantities of PPM in E0.05 and E0.1 prevented sufficient particle 440

    coverage to enable stabilization of the large surface area of such high fraction of oil droplets 441

    (20 wt%). The network formed by PPM in continuous phase appears to encapsulate the oil 442

    droplets, and this prevented the macroscopic oil phase separation (Figure 7). 443

    444

    The remaining three emulsions (E0.25, E0.5 and E1.0) did not show any coalescence upon 445

    dilution within the first week of storage (Supplementary Figure S3) and the ζ-potential values 446

    ranging from -38 to -41 mV (Table 1) indicated that the droplets had high net negative surface 447

    charge providing sufficient electrostatic stability to the droplets. After two weeks of storage, 448

    emulsions E0.25 and E0.5 showed an increase in droplet size (p < 0.05), however, no change 449 in droplet size was observed in the case of E1.0 (p > 0.05) (Table 1). This was associated with 450 decreasing values of ζ-potential for E0.25 and E0.5 (p < 0.05) and no significant change in ζ-451

    potential value in the case of E1.0 (p > 0.05) over time (Table 1). More importantly, after 452 storage over a month, oil droplets were detected in the diluted E0.25 and E0.5 emulsions 453

  • 20

    (Supplementary Figure S3). There was no significant difference in droplet size and ζ-454

    potential of emulsion E1.0 over this 28 day storage period (p> 0.05), which is also in line with 455 results of other Pickering emulsions, where 1 wt% of particles were needed to provide 456

    sufficient surface coverage of the droplets (Araiza-Calahorra, et al., 2019; Du Le, et al., 2020). 457

    In summary, 1.0 wt% PPM was sufficient to create small-sized (d43 ~25 μm) stable droplets 458

    that showed excellent resilience against coalescence (E1.0). Hence, hereafter, 1 wt% PPM was 459

    chosen as the optimum concentration to create Pickering emulsions (E1.0) and test their 460

    response to pH and ionic strengths. 461

    462

    3.3.2 Composition of proteins in the adsorbed phase of E1.0 463

    Compositions of PPC, PPM dispersions and the adsorbed phase of E1.0 were analysed by 464

    reduced SDS-PAGE to understand if there was any difference in composition of the protein 465

    subunits in bulk phase and those that were adsorbed at the interface in case of E1.0. The protein 466

    mixture in PPC or the laboratory-synthesized PPM had more than ten polypeptides (Figure 6) 467

    and the bands at range of 36-42 kDa had the highest proportion. The polypeptide composition 468

    of PPC or PPM was similar to those reported by Mession et al. (2015) and Peng, et al. (2016) 469

    containing the three main proteins, legumin, vicilin and convicilin. Convicilin was the band at 470

    66 kDa. Legumin is a globular protein with an acidic subunit (Lα) at about 42 kDa and basic 471

    subunit (Lβ) at 18- 24 kDa. Vicilin proteins consisted of three subunits (Vi1-3), which are 472

    respectively observed in fractions of around 50 kDa, 36-30 kDa and 20 kDa. Overall, SDS-473

    PAGE results indicate that the formation of PPM from PPC involved all protein subunits, which 474

    is in accordance with results obtained previously that heat treatment did not affect the protein 475

    composition (Laguna, et al., 2017). In addition, the adsorbed phase of E1.0 showed similar 476

    molecular weight profiles to that of PPM that were not influenced by pH-treatment or exposure 477

    to ions. This suggests that adsorption or environmental stresses (pH or ions) applied to PPM 478

  • 21

    had limited effect on the protein composition of the particles. In addition, a significant 479

    proportion of PPM did not enter the resolving gel and were retained in the stacking gel suggest 480

    that they were oligomers above 250 kDa. This suggests that DTT was not sufficient to break 481

    covalent disulphide bridges in the PPM effectively. 482

    483

    3.3.3 Influence of pH on behaviour of E1.0 droplets 484

    Figure 8a shows the mean droplet size distribution of E1.0 (20 wt% sunflower oil) as a function 485

    of pH (pH 3.0, pH 5.0 and pH 7.0) with corresponding Sauter mean diameter (d32), De 486

    Brouckere mean diameter (d43) and ζ-potential shown in Table 2. The initial E1.0 (pH 7.0) has 487

    a significantly larger proportion of droplets in the peak area of 10-100 μm. The smaller peak 488

    in the range of 0.1-1 μm as observed in Figure 8a, overlaps neatly with the size distribution of 489

    PPM estimated using dynamic light scattering (Figure 2), suggesting that these small particles 490

    in Figure 8a might be the free microgel particles that were not adsorbed to the droplet surface 491

    during the homogenization process (Sarkar, et al., 2016b). Comparing the mean diameter of 492

    the PPM particles of ~230 nm (Figure 3a) and the mean size of the oil droplets (d43) of ~ 25 493

    μm (Table 2), the ratio of oil droplet size to PPM size ranges from 100:1 to 1000:1, which is a 494

    signature of a classical Pickering emulsion (Ettelaie, et al., 2015; Sarkar, et al., 2016b). The ζ-495

    potential of E1.0 is about -41 mV (Table 2), similar to that of PPM (Figure 3b) at pH 7.0. This 496

    suggests that perhaps a monolayer of PPM might be present at the surface of droplets. 497

    498

    Interestingly, after adjusting the pH of E1.0 to pH 5.0 or pH 3.0, the d43 values were comparable 499

    to that of E1.0 at neutral pH (p > 0.05). However, d32 values of E1.0 at different pHs showed 500

    a significant difference (p < 0.05) (Table 2). This might be attributed to the fact that d32 value 501

    was more affected by the changes in free (unadsorbed) microgel peak as a function of pH as 502

  • 22

    compared to d43 value, which is line with the behaviour of PPM in aqueous phase as observed 503

    in Figure 3a. The width of the peak at the range 10-100 μm in the droplet size distribution was 504

    narrower with a higher peak height at a lower pH (pH 3.0) than that seen for E1.0 at pH 7.0 505

    (Figure 8a). Also, one might not expect such narrow droplet size distribution in E1.0, 506

    particularly at pH 5.0 considering it is the isoelectric point (pI) of PPM, where E1.0 droplets 507

    will also possess negligible surface charge (Table 2). This is a unique behaviour, unlike that 508

    observed in PPM in the aqueous phase (Figures 3a and 3c) as well as previous studies, where 509

    pH adjustment of emulsions stabilized by pea protein isolate to lower pH increased the 510

    emulsion droplet size and reduced the emulsion stability (Adebiyi, et al., 2011). 511

    512

    To understand the aggregation behaviour, the apparent viscosities of the emulsions (E1.0) at 513

    pH 7.0, pH 5.0 and pH 3.0 were determined using shear rate ranging from 0.1 to 1000 s−1 514

    (Figure 8b). The Ostwald de Waele model was applied to fit the flow curves and the 515

    corresponding fit parameters (consistency coefficient (K), flow index (n), regression coefficient 516

    (R2) were summarized in Table 3. The R2 of all samples was 0.98, confirming a good fit to 517

    the model. For E1.0 at different pH, n varied from 0.63 to 0.80, suggesting that E1.0 was a 518

    pseudoplastic fluid showing shear-thinning behaviour at all the tested pH values. Emulsions at 519

    pH 5.0 showed the highest K and the lowest n (p 0 .05) 525

    in either K or n as compared to those at pH 7.0 (p > 0.05) (Table 3) suggesting that the droplet 526

    flocs that were broken down in the direction of shear were similar at pH 7.0 and pH 3.0. 527

  • 23

    Visual images of E1.0 after 3 months of storage showed no distinct oil layers again confirming 528

    the ability of PPM to to act as effective Pickering stabilizer (Figure 9a). Confocal laser 529

    scanning two-dimensional (2D) (Figure 9a) and three-dimensional (3D) (Figure 9b) 530

    micrographs revealed a thick layer of PPM (Nile Blue staining the protein microgels, displayed 531

    in green) adsorbed at oil-water interface (Nile Red staining the oil droplets, displayed in red), 532

    acting as a barrier to coalescence as observed visually. The confocal micrographs showed 533

    evidence of bridging flocculation as pH was reduced to pI (pH 5.0) with visual signs of 534

    creaming. This is in agreement with the higher viscosities and consequently higher consistency 535

    coefficients of the emulsions at pH 5.0 as compared to those at pH 7.0 and pH 3.0 (p < 0.05) 536

    (Figure 8b, Table 3). When the pH was adjusted to pH 3.0 (Figure 9b), the aggregation of 537

    adsorbed PPM at interface as well as bridging flocculation between the droplets were still 538

    evident. To understand whether the reduction of pH had an effect on PPM that were present at 539

    the interface, Table 2 shows the adsorption efficiency of the PPM as a function of pH. The 540

    PPM had very high degree of adsorption (> 98%) to the droplet surface at all pH (Table 2), 541

    with slightly yet significantly higher adsorption at pH 5.0 as compared to those in pH 7.0 or 542

    pH 3.0 (p < 0.05). This is also evident visually from the images of the subnantant 543

    (Supplementary Figure S4a) after dilution and centrifugation of the emulsions suggesting 544

    that the majority of the PPM particles were either adsorbed at the droplet surface or somehow 545

    associated with interconnecting the neighbouring droplets in a PPM-PPM network. Overall, 546

    E1.0 maintained high stability to coalescence when the pH was adjusted from pH 7.0 down to 547

    pH 5.0 or pH 3.0, where the adsorbed PPM on the droplet surface increased aggregation as 548

    well as the PPM attached to neighbouring droplets. 549

    550

    3.3.4 Influence of background electrolyte concentration on behaviour of E1.0 droplets 551

  • 24

    With the increase in ionic strength from 1 mM to 10 mM, the droplet size distribution (Figure 552

    10a) and corresponding mean droplet diameters (d43, d32) (Table 2), showed no statistically 553

    significant differences (p > 0.05). Although the net surface potential of charged droplets 554

    became less negative (from - 40 mV to -28 mV), the electrostatic repulsion was still sufficient 555

    to inhibit extensive flocculation in E1.0. The emulsions showed shear thinning behaviour 556

    irrespective of ionic strengths (Figure 10b). There was no significant difference between 557

    viscosities and the n values of these emulsions in the presence of 1 mM and 10 mM NaCl, 558

    especially in the region of 0.1-10 s-1 shear rate (p > 0.05) (Table 3). Interestingly, the viscosity 559

    of emulsions in presence of 100 mM NaCl was higher than the other emulsions with 560

    consequently lower n value and higher K value (p < 0.05) (Table 3). The unaltered adsorption 561

    efficiency (Table 2, Supplementary Figure S4b) upon ion treatment (100 mM NaCl) and 562

    enhanced viscosity suggested inter-droplet flocculation in E1.0 with 100 mM NaCl, in line with 563

    lower net surface charge at the droplet surface (-12 mV) (Table 2). No oiling off or phase 564

    separation were observed after rheology measurements, and even after an extended period of 565

    storage (Figure 11). Looking closely at confocal micrographs (Figure 11), droplet flocculation 566

    can be observed after addition of 100 mM NaCl corroborating with the bulk rheological 567

    measurements (Figure 10b). In a previous study, Pickering emulsions stabilized by kafirin 568

    nanoparticles showed a reduction in average droplet size on increasing ionic strength from 10 569

    mM to 50 mM, which was mainly attributed to enhanced nanoparticle interaction via 570

    electrostatic screening effects. On the other hand, de Folter, et al. (2012) suggested that 571

    Pickering emulsions stabilized by both positively- and negatively-charged zein particles at very 572

    high ionic strength (1 M) aggregated and exhibited an emulsion–gel phase. Comparing our 573

    results with these afore-mentioned plant protein-based Pickering emulsions, we hypothesize 574

    that a weak gel-like emulsion structure might have been formed which is apparent from the droplet 575

    aggregation observed in the confocal images (Figure 11). However, the structure of the O/W 576

  • 25

    emulsion might not be as strong as that of a ‘true gel’, but may exhibit a small yield stress, which 577

    requires future rheological characterization. 578

    579

    4. Conclusions 580

    Results from our research demonstrate the ability of a new class of plant protein particles i.e. 581

    pea protein microgels created using a facile top down approach to stabilize O/W Pickering 582

    emulsions with ultrastability against coalescence. To understand the characteristics of these 583

    Pickering O/W emulsion droplets as a function of microgel concentration, pH- or salt-584

    treatment, the colloidal behaviour of pea protein microgel dispersions in aqueous phase was 585

    first investigated at various pH values (pH 2.0 to pH 9.0) or salt concentrations (1 to 250 mM 586

    NaCl) was investigated. Aqueous dispersions of pea protein microgels showed highest degree 587

    of particle aggregation at pH 5.0 as the activation energy barrier in particle-particle interaction 588

    potential was calculated to be extremely low at this pH. Meanwhile, high salt concentrations 589

    resulted in charge screening effects in PPM dispersion but the resulting reduction in 590

    electrostatic potential did not affect the hydrodynamic diameter of microgels, suggesting that 591

    other, possibly steric effects might also be playing a role in the colloidal stability of these 592

    particles. Interestingly, when the pea protein microgels were present at the oil-water interface, 593

    ultra-stable emulsion droplets were obtained only at microgels with 1.0 wt% protein 594

    concentration with all protein subunits i.e. legumin, vicillins and convicillins, being 595

    simultaneously present on the microgel-laden interface. The packed layer of PPM particles 596

    stabilizing the oil droplets allowed the emulsions to be stable over few months against 597

    coalescence. Upon pH reduction to pH 5.0, both intra-droplet and inter-droplet aggregation of 598

    PPM occurred resulting in higher adsorption efficiency and higher viscosity, respectively. The 599

    emulsions also showed responsiveness to ions, especially at 100 mM NaCl with enhancement 600

  • 26

    in viscosity and shear-thinning character. To our knowledge, this is the first comprehensive 601

    study that has systematically demonstrated the role of electrostatics in the colloidal stability of 602

    the plant-based microgel particles in bulk phase versus particles adsorbed at the surface of the 603

    droplets. Findings from this comprehensive study might open door for applicability of these 604

    pea protein-based microgels in a range of food products and allied soft-matter applications, 605

    where alternative plant-based sustainable Pickering stabilizers are increasingly necessary. 606

    Ongoing research is focussing on tuning the surface and bulk properties of pH and ion-607

    responsive pea protein microgel-stabilized emulsions, which can help to tailor the in vitro 608

    gastrointestinal digestion kinetics of these microgel-stabilized emulsified lipids and allow their 609

    usage for controlled release applications in the future. 610

    611

    Acknowledgements 612

    The authors would like to gratefully acknowledge the contributions of Dr Sally Boxal for her 613

    technical support in confocal microscopy at the Bio-imaging Facility within the Faculty of 614

    Biological Sciences of University of Leeds. 615

    References 616

    Adal, Sadeghpour, Connell, Rappolt, Ibanoglu, & Sarkar. (2017). Heteroprotein 617 Complex Formation of Bovine Lactoferrin and Pea Protein Isolate: A 618 Multiscale Structural Analysis. Biomacromolecules, 18(2), 625-635. 619

    Adebiyi, & Aluko. (2011). Functional properties of protein fractions obtained from 620 commercial yellow field pea (Pisum sativum L.) seed protein isolate. Food 621 Chemistry, 128(4), 902-908. 622

    Araiza-Calahorra, & Sarkar. (2019). Pickering emulsion stabilized by protein nanogel 623 particles for delivery of curcumin: Effects of pH and ionic strength on curcumin 624 retention. Food Structure, 21, 100113. 625

    Aveyard, Binks, & Clint. (2003). Emulsions stabilised solely by colloidal particles. 626 Advances in Colloid and Interface Science, 100, 503-546. 627

    Ben-Harb, Panouillé, Huc-Mathis, Moulin, Saint-Eve, Irlinger, Bonnarme, Michon, & 628 Souchon. (2018). The rheological and microstructural properties of pea, milk, 629

  • 27

    mixed pea/milk gels and gelled emulsions designed by thermal, acid, and 630 enzyme treatments. Food Hydrocolloids, 77, 75-84. 631

    Binks. (2002). Particles as surfactants—similarities and differences. Current opinion 632 in colloid & interface science, 7(1-2), 21-41. 633

    Bora, Brekke, & Powers. (1994). Heat Induced Gelation of Pea (Pisum sativum) 634 Mixed Globulins, Vicilin and Legumin. Journal of Food Science, 59(3), 594-635 596. 636

    Chen, Lin, Sun, & Zhao. (2014). Stable and pH-Sensitive Protein Nanogels Made by 637 Self-Assembly of Heat Denatured Soy Protein. Journal of agricultural and food 638 chemistry, 62(39), 9553-9561. 639

    Chihi, Mession, Sok, & Saurel. (2016). Heat-Induced Soluble Protein Aggregates 640 from Mixed Pea Globulins and β-Lactoglobulin. Journal of agricultural and 641 food chemistry, 64(13), 2780-2791. 642

    Chihi, Nicolas, & Saurel. (2018). Acid gelation of mixed thermal aggregates of pea 643 globulins and β-lactoglobulin (Vol. 85). 644

    Cochereau, Nicolai, Chassenieux, & Silva. (2019). Mechanism of the spontaneous 645 formation of plant protein microcapsules in aqueous solution. Colloids and 646 Surfaces A: Physicochemical and Engineering Aspects, 562, 213-219. 647

    Cosgrove. (2010). Colloid science: principles, methods and applications (2nd ed.). 648 Chichester, U.K: Wiley. 649

    David-Birman, Mackie, & Lesmes. (2013). Impact of dietary fibers on the properties 650 and proteolytic digestibility of lactoferrin nano-particles. Food Hydrocolloids, 651 31(1), 33-41. 652

    de Folter, van Ruijven, & Velikov. (2012). Oil-in-water Pickering emulsions stabilized 653 by colloidal particles from the water-insoluble protein zein. Soft Matter, 8(25), 654 6807-6815. 655

    Destribats, Rouvet, Gehin-Delval, Schmitt, & Binks. (2014). Emulsions stabilised by 656 whey protein microgel particles: towards food-grade Pickering emulsions. Soft 657 Matter, 10(36), 6941-6954. 658

    Dickinson. (2012). Use of nanoparticles and microparticles in the formation and 659 stabilization of food emulsions. Trends in Food Science & Technology, 24(1), 660 4-12. 661

    Dickinson. (2013). Stabilising emulsion-based colloidal structures with mixed food 662 ingredients. Journal of the Science of Food and Agriculture, 93(4), 710-721. 663

    Du Le, Loveday, Singh, & Sarkar. (2020). Pickering emulsions stabilised by 664 hydrophobically modified cellulose nanocrystals: Responsiveness to pH and 665 ionic strength. Food Hydrocolloids, 99, 105344. 666

    Ettelaie, & Lishchuk. (2015). Detachment force of particles from fluid droplets. Soft 667 Matter, 11(21), 4251-4265. 668

    Gal, Shani Levi, Tal, & Lesmes. (2013). Emulsions stabilization by lactoferrin nano-669 particles under in vitro digestion conditions (Vol. 33). 670

    Hogg, Healy, & Fuerstenau. (1966). Mutual coagulation of colloidal dispersions. 671 Transactions of the Faraday Society, 62, 1638-1651. 672

  • 28

    Hu, Yin, Zhu, Qi, Guo, Wu, Tang, & Yang. (2016). Fabrication and characterization 673 of novel Pickering emulsions and Pickering high internal emulsions stabilized 674 by gliadin colloidal particles. Food Hydrocolloids, 61, 300-310. 675

    Jayasena, Chih, & Nasar-Abbas. (2010). Functional properties of sweet lupin protein 676 isolated and tested at various pH levels. Research Journal of Agriculture and 677 Biological Sciences, 6(2), 130-137. 678

    Jiao, Shi, Wang, & Binks. (2018). High-Internal-Phase Pickering Emulsions 679 Stabilized Solely by Peanut-Protein-Isolate Microgel Particles with Multiple 680 Potential Applications. Angewandte Chemie International Edition, 57(30), 681 9274-9278. 682

    Karaca, Low, & Nickerson. (2011). Emulsifying properties of chickpea, faba bean, 683 lentil and pea proteins produced by isoelectric precipitation and salt 684 extraction. Food Research International, 44(9), 2742-2750. 685

    Kimura, Fukuda, Zhang, Motoyama, Maruyama, & Utsumi. (2008). Comparison of 686 Physicochemical Properties of 7S and 11S Globulins from Pea, Fava Bean, 687

    Cowpea, and French Bean with Those of Soybean French Bean 7S 688

    Globulin Exhibits Excellent Properties. Journal of agricultural and food 689 chemistry, 56(21), 10273-10279. 690

    Laguna, Picouet, Guàrdia, Renard, & Sarkar. (2017). In vitro gastrointestinal 691 digestion of pea protein isolate as a function of pH, food matrices, 692 autoclaving, high-pressure and re-heat treatments. LWT, 84, 511-519. 693

    Liang, & Tang. (2014). Pea protein exhibits a novel Pickering stabilization for oil-in-694 water emulsions at pH 3.0 (Vol. 58). 695

    Liu, Ou, & Tang. (2017). Ca 2+ -induced soy protein nanoparticles as pickering 696 stabilizers: Fabrication and characterization. Food Hydrocolloids, 65, 175-186. 697

    Liu, & Tang. (2013). Soy protein nanoparticle aggregates as Pickering stabilizers for 698 oil-in-water emulsions. Journal of agricultural and food chemistry, 61(37), 699 8888-8898. 700

    Liu, & Tang. (2014). Emulsifying properties of soy protein nanoparticles: influence of 701 the protein concentration and/or emulsification process. J Agric Food Chem, 702 62(12), 2644-2654. 703

    Liu, & Tang. (2016). Soy glycinin as food-grade Pickering stabilizers: Part. III. 704 Fabrication of gel-like emulsions and their potential as sustained-release 705 delivery systems for β-carotene. Food Hydrocolloids, 56, 434-444. 706

    Makri, & Doxastakis. (2006). Study of emulsions stabilized with Phaseolus vulgaris 707 or Phaseolus coccineus with the addition of Arabic gum, locust bean gum and 708 xanthan gum. Food Hydrocolloids, 20(8), 1141-1152. 709

    Meshulam, & Lesmes. (2014). Responsiveness of emulsions stabilized by lactoferrin 710 nano-particles to simulated intestinal conditions. Food & Function, 5(1), 65-73. 711

    Mession, Chihi, Sok, & Saurel. (2015). Effect of globular pea proteins fractionation 712 on their heat-induced aggregation and acid cold-set gelation. 46, 233-243. 713

    Mession, Roustel, & Saurel. (2017). Interactions in casein micelle – Pea protein 714 system (part I): Heat-induced denaturation and aggregation. Food 715 Hydrocolloids, 67, 229-242. 716

  • 29

    Nesterenko, Alric, Silvestre, & Durrieu. (2013). Vegetable proteins in 717 microencapsulation: A review of recent interventions and their effectiveness. 718 Industrial Crops and Products, 42, 469-479. 719

    Pan, Tikekar, Wang, Avena-Bustillos, & Nitin. (2015). Effect of barrier properties of 720 zein colloidal particles and oil-in-water emulsions on oxidative stability of 721 encapsulated bioactive compounds. Food Hydrocolloids, 43, 82-90. 722

    Peng, Kong, Chen, Zhang, Yang, & Hua. (2016). Effects of heat treatment on the 723 emulsifying properties of pea proteins. Food Hydrocolloids, 52, 301-310. 724

    Peng, Xu, Li, & Tang. (2020). Improving the emulsification of soy β-conglycinin by 725 alcohol-induced aggregation. Food Hydrocolloids, 98, 105307. 726

    Peng, Xu, Liu, & Tang. (2018). Molecular Mechanism for Improving Emulsification 727 Efficiency of Soy Glycinin by Glycation with Soy Soluble Polysaccharide. J 728 Agric Food Chem, 66(46), 12316-12326. 729

    Rayner, Marku, Eriksson, Sjöö, Dejmek, & Wahlgren. (2014). Biomass-based 730 particles for the formulation of Pickering type emulsions in food and topical 731 applications. Colloids and Surfaces A: Physicochemical and Engineering 732 Aspects, 458, 48-62. 733

    Sarkar, Ademuyiwa, Stubley, Esa, Goycoolea, Qin, Gonzalez, & Olvera. (2018). 734 Pickering emulsions co-stabilized by composite protein/ polysaccharide 735 particle-particle interfaces: Impact on in vitro gastric stability. Food 736 Hydrocolloids, 84, 282-291. 737

    Sarkar, Kamaruddin, Bentley, & Wang. (2016a). Emulsion stabilization by tomato 738 seed protein isolate: Influence of pH, ionic strength and thermal treatment. 739 Food Hydrocolloids, 57, 160-168. 740

    Sarkar, Murray, Holmes, Ettelaie, Abdalla, & Yang. (2016b). In vitro digestion of 741 Pickering emulsions stabilized by soft whey protein microgel particles: 742 influence of thermal treatment. Soft Matter, 12(15), 3558-3569. 743

    Sarkar, Zhang, Holmes, & Ettelaie. (2019). Colloidal aspects of digestion of Pickering 744 emulsions: Experiments and theoretical models of lipid digestion kinetics. 745 Advances in Colloid and Interface Science, 263, 195-211. 746

    Shao, & Tang. (2015). Gel-like pea protein Pickering emulsions at pH3.0 as a 747 potential intestine-targeted and sustained-release delivery system for β-748 carotene (Vol. 79). 749

    Tuinier, & De Kruif. (2002). Stability of casein micelles in milk. The Journal of 750 chemical physics, 117(3), 1290-1295. 751

    Xiao, Wang, Gonzalez, & Huang. (2016). Kafirin nanoparticles-stabilized Pickering 752 emulsions: Microstructure and rheological behavior. Food Hydrocolloids, 54, 753 30-39. 754

    Zhu, Zhang, Lin, & Tang. (2017). Freeze-thaw stability of pickering emulsions 755 stabilized by soy and whey protein particles. Food Hydrocolloids, 69, 173-184. 756

    Zhu, Zheng, Liu, Qiu, Lin, & Tang. (2018). Freeze-thaw stability of Pickering 757 emulsions stabilized by soy protein nanoparticles. Influence of ionic strength 758 before or after emulsification. Food Hydrocolloids, 74, 37-45. 759

    760